16
Australian paleo-stress fields and tectonic reactivation over the past 100 Myr R. D. MU ¨ LLER 1 *, S. DYKSTERHUIS 1,2 AND P. REY 1 1 EarthByte Group, School of Geosciences, The University of Sydney, Madsen Building F09, NSW 2006, Australia. 2 ExxonMobil, 12 Riverside Quay, Southbank, VIC 3006, Australia. Even though a multitude of observations suggest time-dependent regional tectonic reactivation of the Australian Plate, its large-scale intraplate stress field evolution remains largely unexplored. This arises because intraplate paleo-stress models are difficult to construct, and that observations of tectonic reactivation are often hard to date. However, because the Australian plate has undergone significant changes in plate boundary types and geometries since the Cretaceous, we argue that even simple models can provide some insights into the nature and timing of crustal reactivation through time. We present Australian intraplate stress models for key times from the Early Cretaceous to the present, and link them to geological observations for evaluating time-dependent fault reactivation. We focus on the effect time-dependent geometries of mid-ocean ridges, subduction zones and collisional plate boundaries around Australia have on basin evolution and fault reactivation through time by reconstructing tectonic plates, restoring plate boundary configurations, and modelling the effect of selected time-dependent plate driving forces on the intraplate stress field of a rheologically heterogeneous plate. We compare mapped fault reactivation histories with paleo-stress models via time-dependent fault slip tendency analysis employing Coulomb-Navier criteria to determine the likelihood of strain in a body of rock being accommodated by sliding along pre-existing planes of weakness. This allows us to reconstruct the dominant regional deformation regime (reverse, normal or strike-slip) through time. Our models illustrate how the complex interplay between juxtaposed weak and strong geological plate elements and changes in far-field plate boundary forces have caused intraplate orogenesis and/or tectonic reactivation in basins and fold belts throughout Australia. KEY WORDS: paleo-stress, fault reactivation, plate tectonics, crustal deformation. INTRODUCTION Faults are a key aspect in many of Australia’s sedimen- tary basins that have undergone significant reactivation resulting in the breaching of hydrocarbon traps. In particular, Oligocene collisional processes north of Australia (Cloetingh et al. 1992) and the Miocene separation of the Indo-Australian Plate into two distinct plates along a diffuse plate boundary (Royer & Chang 1991) have played an instrumental role in modifying the intra-plate stress field, resulting for instance in Miocene reactivation in the Timor Sea (O’Brian et al. 1996). The timing of Cenozoic tectonic events on the Northwest Shelf (NWS) compiled by Cloetingh et al. (1992) indicates a fundamental connection between plate tectonics, i.e. changes in plate motions and related in-plane stresses, and basin subsidence/uplift. As many basins of the Australian Northwest Shelf area have been subjected to a number of repeated extensional and compressional tectonic stages, controlling hydrocarbon migration in fault-controlled traps, understanding fault-trap char- ging and integrity requires some knowledge of paleo- stresses. Etheridge et al. (1991) pointed out that, in particular, steeply dipping strike-slip faults develop into wrench-reactivated transfer faults with associated structures that dominate traps in the Carnarvon, Bonaparte and Gippsland Basins. In central and eastern Australia, major reactivation of basin forming struc- tures occurred in the early Late Cretaceous (ca 95 Ma, e.g. Hill 1994; Korsch et al. 2009) when plate motion to the east slowed down before to change towards a northerly direction. While modelling of the contemporary maximum horizontal stress (s H ) regime is useful for improving our understanding of the driving forces of plate tectonics as well as for the planning of deviated drilling during hydrocarbon production, information concern- ing paleo-stress regimes allows for the creation of predictive frameworks for fault reactivation through time. Compilation of stress data under the auspices of the World Stress Map project in the early 1980s (Zoback 1992) showed s H orientations over most continental areas are parallel to the direction of absolute plate motion, leading to the hypothesis that s H orientations are the product of dominant plate driving forces acting along plate boundaries (Zoback 1992). Orientations of s H 5 10 15 20 25 30 35 40 45 50 55 60 65 70 75 80 85 90 95 100 105 110 115 120 *Corresponding author: [email protected] padmavathym 27/7/11 12:32 TAJE_A_605801 (XML) Australian Journal of Earth Sciences (2011) 00, (1–16) ISSN 0812-0099 print/ISSN 1440-0952 online Ó 2011 Geological Society of Australia DOI: 10.1080/08120099.2011.605801

Australian paleo-stress fields and tectonic …...strong geological plate elements and changes in far-field plate boundary forces have caused intraplate orogenesis and/or tectonic

  • Upload
    others

  • View
    2

  • Download
    0

Embed Size (px)

Citation preview

Page 1: Australian paleo-stress fields and tectonic …...strong geological plate elements and changes in far-field plate boundary forces have caused intraplate orogenesis and/or tectonic

Australian paleo-stress fields and tectonic reactivationover the past 100 Myr

R. D. MULLER1*, S. DYKSTERHUIS1,2 AND P. REY1

1EarthByte Group, School of Geosciences, The University of Sydney, Madsen Building F09, NSW 2006, Australia.2ExxonMobil, 12 Riverside Quay, Southbank, VIC 3006, Australia.

Even though a multitude of observations suggest time-dependent regional tectonic reactivation of theAustralian Plate, its large-scale intraplate stress field evolution remains largely unexplored. This arisesbecause intraplate paleo-stress models are difficult to construct, and that observations of tectonicreactivation are often hard to date. However, because the Australian plate has undergone significantchanges in plate boundary types and geometries since the Cretaceous, we argue that even simplemodels can provide some insights into the nature and timing of crustal reactivation through time. Wepresent Australian intraplate stress models for key times from the Early Cretaceous to the present, andlink them to geological observations for evaluating time-dependent fault reactivation. We focus on theeffect time-dependent geometries of mid-ocean ridges, subduction zones and collisional plateboundaries around Australia have on basin evolution and fault reactivation through time byreconstructing tectonic plates, restoring plate boundary configurations, and modelling the effect ofselected time-dependent plate driving forces on the intraplate stress field of a rheologicallyheterogeneous plate. We compare mapped fault reactivation histories with paleo-stress models viatime-dependent fault slip tendency analysis employing Coulomb-Navier criteria to determine thelikelihood of strain in a body of rock being accommodated by sliding along pre-existing planes ofweakness. This allows us to reconstruct the dominant regional deformation regime (reverse, normal orstrike-slip) through time. Our models illustrate how the complex interplay between juxtaposed weak andstrong geological plate elements and changes in far-field plate boundary forces have causedintraplate orogenesis and/or tectonic reactivation in basins and fold belts throughout Australia.

KEY WORDS: paleo-stress, fault reactivation, plate tectonics, crustal deformation.

INTRODUCTION

Faults are a key aspect in many of Australia’s sedimen-

tary basins that have undergone significant reactivation

resulting in the breaching of hydrocarbon traps. In

particular, Oligocene collisional processes north of

Australia (Cloetingh et al. 1992) and the Miocene

separation of the Indo-Australian Plate into two distinct

plates along a diffuse plate boundary (Royer & Chang

1991) have played an instrumental role in modifying the

intra-plate stress field, resulting for instance in Miocene

reactivation in the Timor Sea (O’Brian et al. 1996). The

timing of Cenozoic tectonic events on the Northwest

Shelf (NWS) compiled by Cloetingh et al. (1992) indicates

a fundamental connection between plate tectonics, i.e.

changes in plate motions and related in-plane stresses,

and basin subsidence/uplift. As many basins of the

Australian Northwest Shelf area have been subjected to

a number of repeated extensional and compressional

tectonic stages, controlling hydrocarbon migration in

fault-controlled traps, understanding fault-trap char-

ging and integrity requires some knowledge of paleo-

stresses. Etheridge et al. (1991) pointed out that, in

particular, steeply dipping strike-slip faults develop into

wrench-reactivated transfer faults with associated

structures that dominate traps in the Carnarvon,

Bonaparte and Gippsland Basins. In central and eastern

Australia, major reactivation of basin forming struc-

tures occurred in the early Late Cretaceous (ca 95 Ma,

e.g. Hill 1994; Korsch et al. 2009) when plate motion to the

east slowed down before to change towards a northerly

direction.

While modelling of the contemporary maximum

horizontal stress (sH) regime is useful for improving

our understanding of the driving forces of plate

tectonics as well as for the planning of deviated drilling

during hydrocarbon production, information concern-

ing paleo-stress regimes allows for the creation of

predictive frameworks for fault reactivation through

time. Compilation of stress data under the auspices of

the World Stress Map project in the early 1980s (Zoback

1992) showed sH orientations over most continental

areas are parallel to the direction of absolute plate

motion, leading to the hypothesis that sH orientations

are the product of dominant plate driving forces acting

along plate boundaries (Zoback 1992). Orientations of sH

5

10

15

20

25

30

35

40

45

50

55

60

65

70

75

80

85

90

95

100

105

110

115

120*Corresponding author: [email protected]

padmavathym 27/7/11 12:32 TAJE_A_605801 (XML)

Australian Journal of Earth Sciences (2011) 00, (1–16)

ISSN 0812-0099 print/ISSN 1440-0952 online � 2011 Geological Society of Australia

DOI: 10.1080/08120099.2011.605801

Page 2: Australian paleo-stress fields and tectonic …...strong geological plate elements and changes in far-field plate boundary forces have caused intraplate orogenesis and/or tectonic

over various portions of the Indo-Australian Plate,

Eurasia and South America, however, have been found

to be more complex and are not parallel to absolute plate

motion trajectories, indicating that plate boundary

forces alone are not enough to understand intra-plate

stress (Heidbach et al. 2007). Here we use the Australian

paleo-stress model presented by Dyksterhuis & Muller

(2008), utilising provinces within the Australian con-

tinent with differing rigidity, and focus on a comparison

with geological observations from a number of different

basins across the continent, particularly on the North-

west Shelf of Australia. If the main features of these

paleo-stress maps can be validated in those areas where

reliable indicators for tectonic reactivation are present,

then the maps have the potential to be used in a

predictive sense in areas with sparse data.

METHODOLOGY

We map Australian Cenozoic intraplate stress evolution

based on Dyksterhuis & Muller (2008), with one addi-

tional but more uncertain model for the Cretaceous (100

Ma), using a methodology described in detail previously

(Dyksterhuis & Muller 2004, 2008; Dyksterhuis et al.

2005a, b). To understand the effect of time-dependent

geometries of mid-ocean ridges, subduction zones and

collisional plate boundaries on the reactivation of

Australia’s main tectonic elements (Figure 1) through

time, we reconstruct tectonic plates, including ocean

floor which has now entirely vanished, restore plate

boundary configurations (Figure 2), and model the effect

of selected time-dependent plate driving forces on the

intraplate stress field. We create a two dimensional,

elastic, plane stress finite element model of the (Indo-)

Australian Plate that distinguishes cratons, fold belts,

basins (Figure 1), and ocean crust in terms of their

relative differences in mechanical stiffness. The numer-

ical model setup and parameters used can be found in

Dyksterhuis & Muller (2008). Orientations of modelled

present-day maximum horizontal stress directions (sH)

agree well with observed directions contained in the

Australian Stress Map Database, with a mean residual

misfit of *128.In order to compare Australia-wide paleo-stress mod-

els based on those described in Dyksterhuis & Muller

(2008) with observations from key basins, we make use of

slip-tendency diagrams, based on a Matlab script devel-

oped by U. Theune and M. Golke (http://www-geo.

phys.ualberta.ca/*utheune/Slip Tendency/slip_html.

html), using a methodology described in Morris et al.

(1996). Slip Tendency (T) is defined as the ratio of shear (t)

to normal stress (sn) normalised by the coefficient of

friction (m):

T ¼ t=ðm� snÞ

The value of the slip tendency for a fault segment is a

measure for its likelihood to fail under the given stress

field. A small value of T indicates stable faults, whereas

125

130

135

140

145

150

155

160

165

170

175

180

185

190

195

200

205

210

215

220

225

230

235

240

245

Figure 1 Topography/bathymetry

map of Australia with model

rheological provinces shown by

grey outlines (relative strengths

listed in Dyksterhuis & Muller

2008).

colour i

n

print &

online

2 R. D. Muller et al.

Page 3: Australian paleo-stress fields and tectonic …...strong geological plate elements and changes in far-field plate boundary forces have caused intraplate orogenesis and/or tectonic

values close to one indicate critically stressed faults. For

a pre-existing fault to be reactivated, the shear stress (t)

on the fault must exceed a certain critical value (tc)

defined as:

tc � m� sn

where m is the coefficient of friction, and sn is the normal

stress acting on the fault plane.

Based on the orientation and magnitude of the

modelled far-field maximum horizontal stress as well

as the magnitude of the minimum horizontal stress

(Table 1) and fault properties (i.e. coefficient of

friction and cohesion) the Matlab code used here

calculates the slip tendency for all possible fault plane

orientations. The strike of modelled fault planes is

varied between 0 and 3608E, and their dip is varied

between 08 and 908, both increasing in steps of 28, so

that 8326 fault plane orientations are analysed for each

case. The computed slip tendencies are plotted as

coloured points on a stereo net. The user can then

superimpose the orientation of a known, existing fault

or fault system for a given area to assess the

likelihood for reactivation under the modelled stress

regime.

Reactivation and inversion of steeply dipping (�608)faults occurs over the Australian continent at various

times. Analogue modelling work by Brun & Nalpas

(1996) demonstrated that transpressional reactivation

(strike-slip reactivation with a component of reverse

motion) occurs along steeply dipping faults when the

direction of sH is oriented at an angle less than 458 to the

strike of the fault. Regional stress regimes are therefore

more likely to reactivate steeply dipping to vertical pre-

existing weaknesses where they are oblique (less than

458) to the structural fabric of a given region.

RESULTS

Modelled present-day sH directions can be directly

compared with measured sH directions in the Austra-

lian Stress Map database to validate the model. The

validation process for the paleo-stress models is not

straightforward given the sparseness of data for direct

comparison. The most useful information for this

purpose is recorded in the geological history as defor-

mation events through time, via successively reacti-

vated generations of faults (normal, reverse, strike-slip)

and folds. In order to validate our paleo-stress models,

250

255

260

265

270

275

280

285

290

295

300

305

310

315

320

325

330

335

340

345

350

355

360

365

370

Figure 2 Indo-Australian Plate with age-area distribution of the oceanic crust (coloured) and location of plate driving forces

applied to a model of the (Indo-) Australian Plate (A) 100 Ma, (B) 55 Ma, (C) 25 to 11 Ma and (D) 11 to 0 Ma. HYM¼Himalayas

(fixed boundary), SUM¼Sumatra Trench, JAVA¼Java Trench, BA¼Banda Arc, PNG¼Papua New Guinea, SOL¼Solomon

Trench, NH¼New Hebrides, NZ¼New Zealand, SNZ¼Australian/Pacific plate margin south of New Zealand,

AAD¼Australia Antarctic Discordance, MQ¼Macquarie Plate, OJP¼Ontong Java Plateau, CS¼Coral Sea, W-

SEIR¼Western Southeast Indian Ridge, E-SEIR¼Eastern Southeast Indian Ridge WBR¼Wharton Basin Ridge, NG¼New

Guinea Margin, NT¼Northern Trench, TSR¼Tasman Sea Ridge, EGS¼East Gondwanaland Subduction Zone. Note that

force arrows are not drawn to scale. Projections used for transforming points of latitude and longitude along the plate

margins to and from a Cartesian reference frame for use in ABAQUS are listed in Appendix 1.

colour i

n

print &

online

Australian paleo-stress 3

Page 4: Australian paleo-stress fields and tectonic …...strong geological plate elements and changes in far-field plate boundary forces have caused intraplate orogenesis and/or tectonic

we have reviewed the deformation record of numerous

sedimentary basins in the Northwest Australian Shelf,

Gippsland Basin, Flinders Ranges and eastern Australia

to test both the spatial and temporal fit of the modelled

sH regime for the modelled time periods.

Tectonic reactivation history of the NorthwestShelf of Australia

The Northwest Australian Shelf (NWS) (Figure 3) has a

general northeast-trending structural fabric inherited

from Late Permian rifting (Keep et al. 1998). Reactiva-

tion and inversion of older structures as well as

generation of anticlines within the basins of the NWS

occurred through time. The Browse Basin holds two

large Miocene inversion structures, the Lombardina

and Lyner structures, interpreted to be transpressional

anticlines (Figure 4) that continued to grow throughout

the Late Miocene (Keep et al. 1998). Marked reactivation

occurs dominantly on steep faults, with associated

flower-type structures with significant inversion. For-

mation of transpressional anticlines is limited to one or

two major faults with strain strongly partitioned along a

few major faults (Keep et al. 1998).

There is a marked change in structural orientation

on the NWS in the Carnarvon Basin from a northeast

orientation in the north Carnarvon (Dampier sub-basin)

to a north-northeast orientation in the south (Exmouth

Sub-Basin) with significant concentration of structural

inversion (Symonds & Cameron 1977). Faulting in the

Carnarvon Terrace (a geographic term for the portion of

the offshore Carnarvon Basin south of the Cape Range

Transform Fault) during the Middle to Late Cretaceous

is restricted to northeast-trending subvertical faults,

which display a wrench reactivation with left-lateral

sense of movement (indicative of a more north–south

oriented maximum horizontal stress direction) (Baillie

& Jacobson 1995). During the Miocene, with some

activity into the present day, wrench reactivation

of northeast-trending subvertical faults is apparent

(Muller et al. 2002) with a right lateral sense of move-

ment (indicative of a more east–west-oriented maximum

horizontal stress direction). There are numerous Mio-

cene inversion structures in the basin, with strain

mainly partitioned along the Rough Range and Lear-

mouth faults (Longley et al. 2002).

The Exmouth Sub-basin experienced at least two

phases of uplift and erosion during the Cretaceous

and inversion and tilting in the Paleogene–Neogene

(Longley et al. 2002). The Barrow Sub-basin just north-

east, however, did not experience this reactivation. In

the Valanginian (137 Ma) the east–west-trending Ninga-

loo Arch was uplifted in the Exmouth Sub-Basin

(Struckmeyer et al. 1998). Uplift occurred along the

west-northwest-trending Novara Arch in the Santonian

(84 Ma) in the Exmouth Sub-Basin, with intense exten-

sion along north-northeast to south-southwest-trending

normal faults associated with this uplift, reactivating

some Triassic/Jurassic faults (Struckmeyer et al. 1998).

Overview of Southeast Australian DeformationHistory

The Flinders Ranges (Figure 5), along with the Mount

Lofty Ranges, are bounded by north–south to northeast–

southwest-trending, moderately (408–458) dipping fault

scarps that demonstrate deformation during the Neo-

gene (Sandiford 2003; Celerier et al. 2005; Quigley et al.

2006). Exposures of range-bounding fault scarps reflect

reverse reactivation (Sandiford 2003) while faults in the

range interior are dominantly strike-slip movement

(Clark & Leonard 2003). In addition, significant Neo-

gene–Quaternary deformation and exhumation has

occurred to the east in the Gippsland Basin in South

Eastern Australia (Dickinson et al. 2001), located

hundreds of kilometres from the nearest plate margin.

During the Oligocene, broad, low-relief northeast-trend-

ing anticlines were formed (Figure 6), while during the

Miocene reverse reactivation of the east–west-trending

margin normal Rosedale and Foster fault systems

375

380

385

390

395

400

405

410

415

420

425

430

435

440

445

450

455

460

465

470

475

480

485

490

495

Table 1 Magnitude and azimuth of modelled maximum horizontal stress (sH) and minimum horizontal stress (sh) averaged over a

small rectangular area indicated by present-day lat./long. coordinates in the table for different time periods.

Time period Azimuth sH (MPa) sh (MPa) Figure key

NW Shelf (Browse) 118/123/-16/-13 (W/E/S/N)

6–0 Ma 70 56 –54 A4

11–6 Ma 67 50 –57 A3

23–11 Ma 117 69 –45 A2

ca 55 Ma 112 6 1 A1

Flinders 138/140/-34/-30 (W/E/S/N)

6–0 Ma 106 9 2 B4

11–6 Ma 69 6 3 B3

23–11 Ma 118 22 79 B2

ca 55 Ma 85 14 2 B1

Gippsland 146/149/-39/-38 (W/E/S/N)

6–0 Ma 129 8 1 C4

11–6 Ma 90 6 3 C3

23–11 Ma 135 25 2 C2

ca 55 Ma 87 1 –0.1 C1

Magnitudes of modelled sH and sh represent the differential stress from a lithostatic reference state of 20 MPa, a coefficient of friction

of 0.6 and a cohesion of 0 MPa (cohesionless faults). ‘Azimuth’ refers to azimuth of the maximum horizontal stress (sH).

4 R. D. Muller et al.

Page 5: Australian paleo-stress fields and tectonic …...strong geological plate elements and changes in far-field plate boundary forces have caused intraplate orogenesis and/or tectonic

500

505

510

515

520

525

530

535

540

545

550

555

560

565

570

575

580

585

590

595

600

605

610

615

620

Figure 4 Miocene inversion structure over the Lombardina structure, Browse Basin, Northwest Shelf, AGSO seismic line 175/

03 (modified from Struckmeyer et al. 1998, where a detailed interpretation of this structure can be found).

Figure 3 Major elements of the

Northwest Australian Shelf show-

ing main basins and main struc-

tural trends (black lines).

Location of seismic line shown in

Figure 4 indicated by red line.

colour i

n

print &

online

colour i

n

print &

online

Australian paleo-stress 5

Page 6: Australian paleo-stress fields and tectonic …...strong geological plate elements and changes in far-field plate boundary forces have caused intraplate orogenesis and/or tectonic

occurred (Figure 7) with up to 1 km of uplift (Dickinson

et al. 2001).

Early Late Cretaceous Tectonic Inversion inCentral and eastern Australia

In the Cretaceous, eastern Australia, then part of

Gondwana’s Pacific margin, was overriding a long-lived

west dipping subduction zone. A succession of previous

extensional events led to the formation of foreland

sedimentary basins (Figure 1) including the Cooper

and Simpson Basins (Permo-Carboniferous to Triassic),

Bowen and Gunnedah Basins (Early Permian to Middle

Triassic), Eromanga and Surat Basins (Early Jurassic to

Early Cretaceous). These basins were inverted during

contractional events, first in the Late Permian to Middle

Triassic (New England Orogen) and later during the

early Late Cretaceous at ca 95 Ma (Hill 1994; Korsch et al.

2009). During the latter, the switch from extension to

contraction lead to uplift, erosion and deep weathering,

folding and reactivation of northeasterly to northerly

trending planar faults such as the Tamworth thrust

(Glen & Brown 1993; Woodward 1995), the Moonie thrust;

and further north the northeast-dipping Jellinbah

thrust zone (Hobbs 1985; Korsch et al. 2009). Some

reactivation led to the formation of hydrocarbon traps

in a number of Australian Basins including the Cooper

Basin (Alexander & Jensen-Schmidt 1996).

The tectonic inversion in the early Late Cretaceous

can be related to a major plate re-organisation. Paleo-

geographic reconstructions show that from 135 to 100 Ma

ago, East Gondwana moved eastwards (Rey & Muller

2010), slowing between 115 and 100 Ma ago. Tectonic

inversion in central and eastern Australia occurred

when Gondwana stopped its eastward motion around

100 Ma, leading to extensional collapse of the Zealandia

Cordillera orogen built along Gondwana’s Pacific mar-

gin (Tulloch et al. 2006), in turn causing a period of

coeval compression landward of the mountain belt (Rey

& Muller 2010). Much later, some time during the

Oligocene or Early Miocene silcrete caps were folded

in central and eastern Australia (Alley 1998). This event

may be linked to the Australia–India collision.

Modelling Results

CRETACEOUS (CA 100 MA)

In addition to the paleo-stress time slices discussed by

Dyksterhuis & Muller (2008), we present one tentative

model for the Cretaceous. During the Cretaceous,

following Greater India–Australia break-up at ca 132

Ma, the Australian intraplate stress field was dominated

by ridge push from the now subducted northern

Wharton Basin Ridge in the Neo-Tethys Ocean (Figure

2A). Modelled sH directions for the Cretaceous (Figure 8)

generally trend north–south over the entire continent.

However, we regard the details in this model as

625

630

635

640

645

650

655

660

665

670

675

680

685

690

695

700

705

710

715

720

725

730

735

740

745

750

Figure 5 Relief map of the Flinders Ranges and the Gipps-

land Basin showing major faults and locations of seismic

lines.

Figure 6 Seismic line G92A-3125

(see Figure 5 for location) offshore

Gippsland Basin. Early Miocene

sediments onlap (blue arrows) the

deeper structure, indicating defor-

mation within the Oligocene–

Early Miocene. Onlapping reflec-

tors, stratigraphic thinning and

folding higher in the section in-

dicate that deformation continued

into the Plio-Pleistocene. Differ-

ent stratigraphic thicknesses on

either side of the reverse fault

(red arrows) indicate fault move-

ment within the Oligocene–Early

Miocene (modified from Dickin-

son et al. 2001).

colour i

n

print &

online

colour i

n

print &

online

6 R. D. Muller et al.

Page 7: Australian paleo-stress fields and tectonic …...strong geological plate elements and changes in far-field plate boundary forces have caused intraplate orogenesis and/or tectonic

substantially more uncertain than our Cenozoic models,

and mainly include it to demonstrate the overall

difference between the Cretaceous and the Eocene stress

fields. With an average azimuth of roughly 608 to the

main structural fabric in the Browse Basin, the

orientation of sH directions is not oblique enough to

cause any significant reverse reactivation or inversion

of pre-existing northeast-trending structures, though

strike-slip reactivation is possible. We do not regard the

modelled stress directions in central and eastern

Australia in this model as well constrained at this time,

because mantle-driven forces may have played a large

role (Matthews et al. 2011) relative to plate boundary

forces at that time, in modulating the regional paleo-

stress field.

EARLY LATE CRETACEOUS TO EOCENE (CA 95–55 MA)

The Australian stress field is believed to have changed

dramatically in the early Late Cretaceous (ca5 95 Ma)

due to the establishment of plate driving forces from

mid-ocean ridges to the east and south of Australia

(Gaina & Muller 2007) (Figure 2B). Modelled sH

magnitudes over Australia for the Eocene are rela-

tively low at this time (Figure 9), with sH orientations

now trending generally west-northwest over the NWS,

and we propose that this tectonic regime was estab-

lished at ca 95–90 Ma when a mid-ocean ridge was

developing east of Australia. Reactivation of the

northeast-trending structural fabric on the NWS at

this time is most likely in a strike-slip regime (Figure

13A1).2 The northeast-trending modelled sH orienta-

tions on the Exmouth Plateau are compatible with the

formation of northeast-trending anticlines, however,

modelled sH magnitudes at this time are low, suggest-

ing that folding may be unlikely.

Modelled sH directions rotate to a roughly east–

west orientation in the Flinders Ranges (Figure 9),

central and eastern Australia, approximately orthogo-

nal to the general north–south trending pre-existing

fault fabric (Figure 13B1). In the Flinders Ranges, this

is associated with a dramatic change in strain regime

from strike-slip to reverse, while in central and

eastern Australia the new stress regime explains the

tectonic inversion of north–south trending normal

faults. Slip tendency graphs (Figure 13B1) indicate

east–west sH directions to be optimally orientated for

reverse reactivation of the moderately dipping faults

at shallow depths (Sandiford 2003). This suggests an

Early Eocene onset of the Sprigg’s Orogeny via

reactivation of pre-existing faults (Dyksterhuis &

Muller 2008). Evidence from apatite fission track

analysis indicates that regional cooling of up to 608Cand uplift occurred in the Flinders Ranges during the

Eocene (Mitchell et al. 2002), accompanied by an

increase in sedimentation rates into surrounding

basins, supporting our model. Modelled sH magnitudes

in the Gippsland Basin at ca 55 Ma show the region to

be marginally in the strike-slip tectonic stress regime

(Figure 13C1). Modelled stress magnitudes are very

low, however, and other portions of southeast Aus-

tralia appear to be in an extensional tectonic stress

regime (Figure 9). This is supported the observation

that extensional tectonism prevailed in the Gippsland

Basin until the Early Eocene (Johnstone et al. 2001;

Sandiford et al. 2004) with a compressional tectonic

regime not becoming dominant until after the cessa-

tion of seafloor spreading in the Tasman Sea at ca 52

Ma (Bernecker et al. 2006).

A significant basin-forming event occurred along

the east coast of Queensland in the Middle to Late

Eocene, where a series of en-echelon graben and half-

graben basins formed striking north-northwest (Gib-

son 1989). These basins host oil shales which have

been explored since 1980 for hydrocarbons (McConno-

chie & Henstridge 1985). Our Eocene model shows

extremely small-magnitude maximum horizontal stres-

ses (Figure 9), bordering on an extensional regime, in

qualitative agreement with the Eocene change in

stress regime indicated by graben formation in East

Queensland. Our modelled horizontal maximum stress

directions are not oriented parallel to the strike of the

grabens, as would be expected. This may be a

consequence of uncertainties in the distribution of

relatively weak and strong lithospheric elements in

the region.

755

760

765

770

775

780

785

790

795

800

805

810

815

820

825

830

835

840

845

850

855

860

865

870

875

Figure 7 Seismic line G92A-3038

offshore Gippsland (see Figure 5

for location) showing relatively

young (post-Middle Miocene) re-

verse movement on the Rosedale

Fault System (modified from Dick-

inson et al. 2001).

colour i

n

print &

online

Australian paleo-stress 7

Page 8: Australian paleo-stress fields and tectonic …...strong geological plate elements and changes in far-field plate boundary forces have caused intraplate orogenesis and/or tectonic

EARLY–MIDDLE MIOCENE (23–11 MA)

The sH magnitudes increase in amplitude over the NWS

in the Early Miocene model (Figure 10) while sH

orientations remain broadly similar to those during

the Eocene, trending generally northwest on the NWS.

Orientations of modelled sH directions remain north-

west-trending in the Exmouth Plateau area between the

Eocene and Early Miocene, however, the amplitude of

the modelled sH magnitudes increases between the two

time periods, suggesting that the formation of the

northeast-trending antliclines (Symonds et al. 1994)

occurred more likely between ca 55 to ca 23 Ma. This

corresponds to a time period when spreading in the

Tasman Sea ceased at ca 52 Ma, followed by a cessation

of spreading in the Wharton Basin at ca 42 Ma (Krishna

et al. 1995). The hard collision between India and

Eurasia was likely initiated at ca 34 Ma (Aitchison

880

885

890

895

900

905

910

915

920

925

930

935

940

945

950

955

960

965

970

975

980

985

990

995

1000

Figure 8 Modelled maximum horizontal stress for the Cretaceous (ca 100 Ma). Stress magnitude units are in MPa.

colour i

n

print &

online

8 R. D. Muller et al.

Page 9: Australian paleo-stress fields and tectonic …...strong geological plate elements and changes in far-field plate boundary forces have caused intraplate orogenesis and/or tectonic

et al. 2007), paired with the initiation of mountain

building along Australia’s eastern plate boundary along

the Alpine Fault at ca 25 Ma (Kamp & Fitzgerald 1987).

Together, these severely altered boundary forces estab-

lished a stress regime in the Miocene dominated by

ridge push south of Australia and increased resisting

forces north of India and east of Australia, raising

compressive stress magnitudes on the NWS severely in

the Miocene compared with older model times, in

agreement with structural observations indicating

widespread Neogene tectonic reactivation in the Timor

Sea (Shuster et al. 1998).

Modelled sH directions in the Flinders Ranges trend

northwest in the Early to Middle Miocene. Magnitudes

of modelled sH indicate a moderate likelihood of strike-

slip reactivation of pre-existing faults at this time, with

reverse reactivation of range bounding faults less likely

compared with 55 Ma or the present day (Figure 13B2),

1005

1010

1015

1020

1025

1030

1035

1040

1045

1050

1055

1060

1065

1070

1075

1080

1085

1090

1095

1100

1105

1110

1115

1120

1125

Figure 9 Modelled maximum horizontal stress for the Eocene (ca 55 Ma). Stress magnitude units are in MPa.

colour i

n

print &

online

Australian paleo-stress 9

Page 10: Australian paleo-stress fields and tectonic …...strong geological plate elements and changes in far-field plate boundary forces have caused intraplate orogenesis and/or tectonic

suggesting a period of quiescence in the Flinders Ranges

during the Early to Middle Miocene (Dyksterhuis &

Muller 2008).

Modelled sH magnitudes in the Gippsland Basin were

greatest in the Early to Middle Miocene with modelled

sH directions trending southeast with a mean azimuth of

1348. A reverse faulting stress regime was extant and

reactivation of steeply dipping, east–west-trending

faults in the region, is interpreted to be moderately to

highly likely (Figure 13C2). These results are consistent

with observations that show compression along north-

east–southwest-trending anticlines peaking during the

Middle Miocene (Dickinson et al. 2001; Bernecker et al.

2006) with contemporaneous transpressional reac-

tivation of east–west-trending faults (Dickinson et al.

2001).

1130

1135

1140

1145

1150

1155

1160

1165

1170

1175

1180

1185

1190

1195

1200

1205

1210

1215

1220

1225

1230

1235

1240

1245

1250

Figure 10 Modelled maximum horizontal stress for the Early to Middle Miocene (23–11 Ma). Stress magnitude units are in

MPa.

colour i

n

print &

online

10 R. D. Muller et al.

Page 11: Australian paleo-stress fields and tectonic …...strong geological plate elements and changes in far-field plate boundary forces have caused intraplate orogenesis and/or tectonic

The increase in modelled stress magnitudes in much

of the Eromanga Basin in the Early Miocene is likely

related to a period of broad-wavelength folding and

faulting, observed in Cretaceous sediments in north-

eastern South Australia at various localities along the

edge of the Simpson Desert and in the southwest

Queensland portion of the Eromanga Basin (Sprigg

1963; Wopfner 1974). In addition, Gregory et al. (1967)

and Senior et al. (1968) mapped about 40 equivalent

anticlines in the southwest Queensland portion of the

Eromanga Basin, based on the deformation of a

previously formed silcrete surface, marking an origin-

ally flat-lying unconformity (Sprigg 1963). The asso-

ciated fold axes are approximately NE–SW oriented and

the fold wavelengths are up to around 50 km and larger

with amplitudes of about 20 to 200 m (Sprigg 1986).

Mavromatidis (2006) mapped a maximum uplift of 180 m

associated with this deformation period. Finlayson &

1255

1260

1265

1270

1275

1280

1285

1290

1295

1300

1305

1310

1315

1320

1325

1330

1335

1340

1345

1350

1355

1360

1365

1370

1375

1380

Figure 11 Modelled maximum horizontal stress for the Late Miocene (11–6 Ma). Stress magnitude units are in MPa.

colour i

n

print &

online

Australian paleo-stress 11

Page 12: Australian paleo-stress fields and tectonic …...strong geological plate elements and changes in far-field plate boundary forces have caused intraplate orogenesis and/or tectonic

Leven (1988) and Wopfner (1978) estimated the timing of

the deformation as Eogene/Oligocene–Early Miocene, in

broad agreement with our model results.

LATE MIOCENE (11–6 MA)

The modelled sH regime (Figure 11) changes consider-

ably at this time with the onset of collision at the Papua

New Guinea margin (Hall 2002). This oblique collision is

expressed by Miocene–Recent anomalous subsidence in

the Cartier Trough and the Malita Graben, likely related

to reactivation of rift structures by sinistral shear.

Other structures in the Timor Sea have been inverted by

left-lateral shear at confining bends of faults (Shuster

et al. 1998). While previous models showed consistent

trends for sH directions over the entire NWS, in the Late

1385

1390

1395

1400

1405

1410

1415

1420

1425

1430

1435

1440

1445

1450

1455

1460

1465

1470

1475

1480

1485

1490

1495

1500

1505

Figure 12 Modelled contemporary (6–0 Ma) maximum horizontal stress. Orientation data from the Australian Stress Map

Database shown by white filled bars. Stress magnitude units are in MPa.

colour i

n

print &

online

12 R. D. Muller et al.

Page 13: Australian paleo-stress fields and tectonic …...strong geological plate elements and changes in far-field plate boundary forces have caused intraplate orogenesis and/or tectonic

Miocene sH orientations rotate in a clockwise direction

from east–west in the Carnarvon Basin to northeast in

the Timor Sea. The mean sH orientation of 678 in the

Browse Basin at this time indicates that transpressional

reactivation of northeast-trending faults is likely (Fig-

ure 13A3). The documented right-lateral sense of move-

ment on the faults at this time (Etheridge et al. 1991)

agrees well with the general east–west trend of the

modelled sH directions.

Modelled sH orientations in the Flinders Ranges

during the Middle to Late Miocene show increased

variability compared with previous time periods. Mod-

elled sH directions were oriented roughly north–south

over the Mt Lofty Ranges in the southern Adelaide Fold

Belt at this time and roughly east–west over the Flinders

Ranges toward the north. Conditions favourable for

reverse reactivation along the Flinders Ranges were re-

established in the Middle to Late Miocene and reverse

reactivation of moderately dipping, north–south-trend-

ing range bounding faults in the Flinders Ranges was

likely at this time (Figure 13B3), potentially triggering

the second, ongoing phase of the Sprigg’s Orogeny

(Dyksterhuis & Muller 2008). Based on observations

from the Simpson Desert, Sprigg (1963) estimated that

basin (and gentle syncline) formation, coincident with

Simpson Desert depression, has continued during the

Neogene in the southwestern Eromanga Basin. This is

compatible with our model results. Modelled sH direc-

tions in the Gippsland Basin rotate to an east–west

orientation in the Middle to Late Miocene and a reverse

stress regime existed. Reactivation of east–west-trend-

ing, steeply dipping faults is interpreted as unlikely at

this time (Figure 13C3).

CONTEMPORARY (6–0 MA)

Convergence speed along the southern portion of the

Pacific–Australia plate boundary increased twofold at

1510

1515

1520

1525

1530

1535

1540

1545

1550

1555

1560

1565

1570

1575

1580

1585

1590

1595

1600

1605

1610

1615

1620

1625

1630

Figure 13 Equal area lower hemisphere stereonets indicating slip tendency, with hotter colours indicating greater likelihood

slip to occur for particular combinations of dip and strike of a given fault. Uniformly red colours in diagrams A2–4, indicate

differential stress saturation, for the Miocene to present Northwest Shelf result from very large differences in modelled

maximum and minimum horizontal stresses. However, it needs to be kept in mind that our models are simplified, and do not

take into account depth-dependence of fault reactivation, mantle processes or plate flexure—therefore these models become

less robust close to plate boundaries. Especially the modelled minimum horizontal stresses here are likely too low and may

be in error, reflecting oversimplications in our model in regions close to plate boundaries. The dominant stress regime at

these times (strike slip) is nevertheless interpreted to be the most likely regime with faults oriented at *458 to the maximum

horizontal stress orientation, the most favourable orientation for strike-slip reactivation. Planes are represented on the

stereonets by a single point placed at 908 to the strike of the plane with the dip of the plane indicated by the pole’s proximity to

the centre (the large black stippled circle indicates a dip of 608, the small black stippled circle indicates a dip of 308 with the

centre representing a dip of 08). Slip tendency graphs are computed for a lithostatic stress state of 20 MPa, a coefficient of

friction of 0.6 and a cohesion of 0 MPa (cohesionless faults). Slip tendency graphs for three regions at various modelled times

are shown: North West Shelf (Browse Basin) (A1–A4) [(A1) 55 Ma, (A2) 23 to 11 Ma, (A3) 11 to 6 Ma and (A4) 6 to 0 Ma], the

Flinders Ranges (B1–B4) [(B1) 55 Ma, (B2) 23 to 11 Ma, (B3) 11 to 6 Ma and (B4) 6 to 0 Ma] and the Gippsland Basin (C1–C4) [(C1)

55 Ma, (C2) 23 to 11 Ma, (C3) 11 to 6 Ma and (C4) 6 to 0 Ma]. See Table 1 for mean sH azimuth and principle stress magnitudes

for individual time periods. The white line indicates the strike of the structural fabric for the region with the white arrow

indicating the pole to the plane indicating general dip of faults. The modelled maximum horizontal stress orientation is

overlain in pink. Andersonian stress regime style is indicated by text at the lower right of each stereonet where NF¼normal

faulting, TF¼ thrust faulting and SS¼ strike-slip faulting.

colour i

n

print &

online

Australian paleo-stress 13

Page 14: Australian paleo-stress fields and tectonic …...strong geological plate elements and changes in far-field plate boundary forces have caused intraplate orogenesis and/or tectonic

about 6 Ma (Cande & Stock 2004) (Figure 2D). Changes in

compressional forces along the New Zealand margin

between the Middle Miocene and present day do not

greatly effect the modelled sH regime over the NWS,

with present day modelled sH directions and magnitudes

(Figure 12) similar to those in the Middle–Late Miocene.

Transpressional reactivation of northeast-trending stee-

ply dipping faults in the Browse Basin is likely based on

slip tendency analysis (Figure 13A4).

Modelled sH magnitudes over the Flinders Ranges

increase in response to increased convergence along the

southeast margin of the Indo-Australian Plate. Reverse

reactivation of moderately dipping, north–south-trend-

ing range bounding faults in the Flinders Ranges is still

likely at this time (Figure 13B4). High Quaternary slip

rates on range bounding faults showing reverse sense

motion in the Flinders Ranges (Sandiford 2003) demon-

strate evidence for this modelled sH regime.

Modelled present-day sH directions in the Gipps-

land Basin trend northwest and agree with observed

northwest-trending sH orientations (Hillis & Reynolds

2000). Modelled sH directions and magnitudes indicate

that reverse reactivation is likely (Figure 13C4) and

are supported by observations of continued growth on

northeast-trending anticlines in the region into the

Pliocene. Dickinson et al. (2001) observed that up to

one kilometre of exhumation or more has occurred

in the Gippsland and Otway Basins post Middle

Miocene.

CONCLUSIONS

Our models indicate that significant changes in the

stress regime of the geologically heterogeneous Austra-

lian continent through time can be modelled by chan-

ging geometry and forces acting along the boundaries of

the Indo-Australia and Paleo-Australian plate since the

Early Cretaceous and we demonstrate that intraplate

structural events may be caused by interaction of the far

field stress field with the heterogeneous geology of

Australia. We validate our paleo-stress models by

examining the fault reactivation histories of various

sedimentary basins in Australia. Our modelling indi-

cates that intraplate suture zones of the Australian

continent that were particularly weak, such as the

strongly faulted portions of the Northwest Shelf and the

Flinders Ranges, have been reactivated during times

when favourable stress regimes existed. Our modelling

approach, which includes spatial variations in plate

stiffness, clearly supports the hypothesis that pre-

existing weaknesses in continental crust can act to

localise deformation, and we show quantitatively how

juxtaposed strengths and weaknesses in the plate

together with time-varying plate driving forces act to

reactivate particular portions of the continental crust

with tectonic fabric orientations favourable for reacti-

vation through time. Our models provide a tectonic

framework, validated with geological observations, for

understanding first-order structural reactivation of the

Australian continent and shelf. The digital paleo-stress

models discussed in this paper are accessible at http://

www.earthbyte.org.

ACKNOWLEDGEMENTS

We thank Judith Sippel and Alexey Goncharov for their

constructive reviews that improved the paper substan-

tially. RDM was supported by ARC FL0992245 and PR by

ARC grant DP0987604.

REFERENCES

AITCHISON J., ALI J. R. & DAVIS A. M. 2007. When and where did India

and Asia collide? Journal of Geophysical Research 112,

doi:10.1029/2006JB004706.

ALEXANDER E. M. & JENSEN-SCHMIDT B. 1996. Structural and tectonic

history. Department of Mines and Energy. The Petroleum Geology

of South Australia.

ALLEY N. F. 1998. Cainozoic stratigraphy, palaeoenvironments and

geological evolution of the Lake Eyre Basin. Palaeogeography,

Palaeoclimatology, Palaeoecology 144, 239–263.

BAILLIE P. W. & JACOBSON E. 1995. Structural evolution of the

Carnarvon Terrace, Western Australia. The APEA Journal 35,

321–332.

BERNECKER T., THOMAS J. H. & O’BRIEN G. W. 2006. Hydrocarbon

prospectivity of areas V06-2, V06-3 and V06-4, southern offshore

Gippsland Basin, Victoria, Australia. Industries V. D. o. P.

BRUN J. P. & NALPAS T. 1996. Graben inversion in nature and

experiments. Tectonics 15, 677–687.

CANDE S. C. & STOCK J. M. 2004. Pacific–Antarctic–Australia motion

and the formation of the Macquarie plate. Geophysical Journal

International 157, 399–414.

CELERIER J., SANDIFORD M., HANSEN D. L. & QUIGLEY M. 2005. Modes

of active intraplate deformation, Flinders Ranges, Australia.

Tectonics 24, C6006–C6006.

CLARK D. & LEONARD M. 2003. Principal stress orientations from

multiple focal-plane solutions: new insight into the Australian

intraplate stress field. In: Hillis R. R. & Muller D. eds. Evolution

and dynamics of the Australian Plate, Vol. 22, p. 438, Geological

Society of Australia Special Publication, Sydney.

CLOETINGH S., STEIN C., REEMST P., GRADSTEIN F., WILLIAMSON P., EXON

N. & VON RAD U. 1992. Continental margin stratigraphy, deforma-

tion, and intraplate stresses for the Indo-Australian region. In:

Gradstein F. M., Ludden J. N. et al. eds. Proceedings of the Ocean

Drilling Program, scientific results, Vol. 123, pp. 671–713.

DICKINSON J. A., WALLACE M. W., HOLDGATE G., DANIELS J.,

GALLAGHER S. J. & THOMAS L. 2001. Neogene tectonics in SE

Australia: Implications for Petroleum Systems. APPEA Journal

41, 37–51.

DYKSTERHUIS S. & MULLER D. 2004. Modelling the contemporary and

palaeo stress field of Australia using finite-element modelling

with automatic optimisation. Exploration Geophysics 35, 236–241.

DYKSTERHUIS S. & MULLER R. D. 2008. Cause and evolution of

intraplate orogeny in Australia. Geology 36, 495–498.

DYKSTERHUIS S., ALBERT R. A. & MULLER D. 2005a. Finite element

modelling of intraplate stress using ABAQUSTM. Journal of

Computers and Geosciences 31, 297–307.

DYKSTERHUIS S., MULLER R. D. & ALBERT R. A. 2005b. Palaeo-stress

field evolution of the Australian continent since the Eocene.

Journal of Geophysical Research 110, B05102 doi:05110.01029/

02003JB002728.

ETHERIDGE M., MCQUEEN H. & LAMBECK K. 1991. The role of

intraplate stress in Tertiary (and Mesozoic) deformation of the

Australian continent and its margins: A key factor in petroleum

trap formation. Exploration Geophysics 22, 123–128.

FINLAYSON D. M. & LEVEN J. H. 1988. Structural Styles and Basin

Evolution in Eromanga Region, Eastern Australia. American

Association of Petroleum Geologists 72.

GAINA C. & MULLER R. D. 2007. Cenozoic tectonic and depth/age

evolution of the Indonesian gateway and associated back-arc

basins. Earth Science Reviews 83, 177–203.

GIBSON P. J. 1989. Petrology of two Tertiary oil shale deposits from

Queensland, Australia. Journal of the Geological Society of

London 146, 319–331.

3GLEN R. A. & BROWN R. E. 1993. A transect through a forearc basin:

preliminary results from a transect across the Tamworth Belt at

Manilla N.S.W. Department of Geology & Geophysics, University

of New England New England Orogen, eastern Australia.

1635

1640

1645

1650

1655

1660

1665

1670

1675

1680

1685

1690

1695

1700

1705

1710

1715

1720

1725

1730

1735

1740

1745

1750

1755

14 R. D. Muller et al.

Page 15: Australian paleo-stress fields and tectonic …...strong geological plate elements and changes in far-field plate boundary forces have caused intraplate orogenesis and/or tectonic

GREGORY C. M., SENIOR B. R. & GALLOWAY M. C. 1967. The geology of

the Connemara, Jundah, Canterbury, Windorah and Adavale

1:250,000 shet areas, Queensland. Bureau of Mineral Resources.

HALL R. 2002. Cenozoic geological and plate tectonic evolution of SE

Asia and the SW Pacific: computer-based reconstructions, model

and animations. Journal of Asian Earth Sciences 20, 353–431.

HEIDBACH O., REINECKER J., TINGAY M., MULLER B., SPERNER B.,

FUCHS K. & WENZEL F. 2007. Plate boundary forces are not

enough: Second- and third-order stress patterns highlighted in

the World Stress Map database. Tectonics 26, doi:10.1029/

2007TC002133.

HILL P. J. 1994. Geology and geophysics of the offshore Marybor-

ough, Capricorn and northern Tasman Basins. AGSO Australian

Geological Survey Organisation Record.

HILLIS R. R. & REYNOLDS S. D. 2000. The Australian stress map.

Journal of the Geological Society of London 157, 915–921.

HOBBS B. E. 1985. Interpretation and analysis of structure in the

Bowen Basin (unpubl.).

JOHNSTONE R. D., JENKINS C. C. & MOORE M. A. 2001. An integrated

structural and palaeogeographic investigation of Eocene ero-

sional events and related hydrocarbon potential in the Gipps-

land Basin. Eastern Australasian Basins Symposium 2001,

Melbourne, pp. 403–412. The Australasian Institute of Mining

and Metallurgy.

KAMP P. J. J. & FITZGERALD P. G. 1987. Geologic constraints on the

Cenozoic Antarctica–Australia–Pacific relative plate motion

circuit. Geology 15, 694–697.

KEEP M., POWELL C. McA. & BAILLIE P. W. 1998. Neogene Deforma-

tion of the North West Shelf, Australia. The Sedimentary Basins

of Western Australia 2. Perth, Australia, pp. 81–94. Petroleum

Exploration Society of Australia.

KORSCH R. J., TOTTERDELL J. M., FOMIN T. & NICOLL M. G. 2009.

Contractional structures and deformational events in the

Bowen, Gunnedah and Surat Basins, eastern Australia. Austra-

lian Journal of Earth Sciences 56, 477–499.

KRISHNA K. S., RAO D. G., RAMANA M. V., SUBRAHMANYAM V., SARMA

K. V. L. N. S., PILIPENKO A. I., SHCHERBAKOV V. S. and

RADHAKRISHNA MURTHY I. V. 1995. Tectonic model for the

evolution of oceanic crust in the northeastern Indian Ocean

from the Late Cretaceous to the early Tertiary. Journal of

Geophysical Research 100, 20011–20024.1LONGLEY I. M., BUESSENSCHUETT C., CLYDESDALE L., CUBITT C. J.,

DAVIS R. C., JOHNSON M. K., MARSHALL N. M., SOMERVILLE R.,

SPRY T. B. & THOMPSON N. B. 2002. The North West Shelf of

Australia—a Woodside perspective. In: Keep M. & Moss S. J. eds.

The sedimentary basins of Western Australia 3, pp. 27–86,

Petroleum Exploration Society of Australia, Perth, WA.

MATTHEWS K. J., HALE A. J., GURNIS M., MULLER R. D. & DICAPRIO L.

2011. Dynamic subsidence of Eastern Australia during the

Cretaceous. Gondwana Research 19, 372–383.

MAVROMATIDIS A. 2006. Burial/exhumation histories for the Cooper

Eromanga Basins and implications for hydrocarbon exploration,

Eastern Australia. Basin Research 18, 351–373.

MCCONNOCHIE M. J. & HENSTRIDGE D. A. 1985. The Lowmead

Graben—geology, Tertiary oil shale genesis and regional tec-

tonic implications. Australian Journal of Earth Sciences 32, 205–

218.

MITCHELL M. M., KOHN B. P., O’SULLIVAN P. B., HARTLEY M. J. &

FOSTER D. A. 2002. Low-temperature thermochronology of the Mt

Painter Province, South Australia. Australian Journal of Earth

Sciences 49, 551–563.

MORRIS A., FERRILL D. A. & HENDERSON D. B. 1996. Slip-tendency

analysis and fault reactivation. Geology 24, 275–278.

MULLER D., MIHUT D., HEINE C., O’NEIL C. & RUSSELL I. 2002. Tectonic

and volcanic history of the Canarvon Terrace: Constraints from

seismic interpretation and geodynamic modelling. The Sedimen-

tary Basins of Western Australia 3, Perth, Western Australia, pp.

719–740. Petroleum Exploration Society of Australia.

O’BRIAN G. W., LISK M., DUDDY I., EADINGTON P. J., CADMAN S. &

FELLOWS M. 1996. Late Tertiary fluid migration in the Timor Sea:

A key control on thermal and diagenetic histories? APEA

Journal 1996, 399–426.

QUIGLEY M. C., CUPPER M. L. & SANDIFORD M. 2006. Quaternary faults

of south-central Australia: palaeoseismicity, slip rates and

origin. Australian Journal of Earth Sciences 53, 285–301.

REY P. F. & MULLER R. D. 2010. Fragmentation of active continental

margins owing to the buoyancy of the mantle wedge. Nature

Geoscience 3, 257–261.

ROYER J-Y. & CHANG T. 1991. Evidence for relative motions between

the Indian and Australian plates during the last 20 m.y. from

plate tectonic reconstructions: implications for the deformation

of the Indo-Australian plate. Journal of Geophysical Research 96,

11779–11802.

SANDIFORD M. 2003. Neotectonics of southeastern Australia: linking

the Quaternary faulting record with seismicity and in situ

stress. In: Hillis R. R. & Muller D. eds., Evolution and dynamics of

the Australian Plate, Vol. 22, p. 438, Geological Society of

Australia Special Publication, Sydney.

SANDIFORD M., WALLACE M. & COBLENTZ D. 2004. Origin of the in situ

stress field in south-eastern Australia. Basin Research 16, 325–

338.

SENIOR B. R., GALLOWAY M. C., INGRAM J. A. & SENIOR D. 1968. The

geology of the Barrolka, Eromanga, Durham Downs, Thargo-

mindah, Tickalara and Bulloo 1:250,000 Sheet areas, Queensland.

Bureau of Mineral Resources.

SHUSTER M. W., EATON S., WAKEFIELD L. L. & KLOOSTERMAN H. J.

1998. Neogene tectonics, greater Timor Sea, offshore Australia:

Implications for trap risk. The APPEA Journal 38, 351–379.

SPRIGG R. C. 1963. Geology and Petroleum prospects of the Simpson

Desert. Transactions of the Royal Society of South Australia 86,

36–65.

SPRIGG R. C. 1986. The Eromanga Basin in the search for commercial

hydrocarbons. In: Gravestock D. I., Moore P. S. & Pitt G. M. eds.,

Contributions to the geology and hydrocarbon potential of the

Eromanga Basin, Vol. Special Publ. 12, pp. 9–24, Geol. Soc. Aust.,

Sydney.

STRUCKMEYER H. I. M., BLEVIN J. E., SAYERS J., TOTTERDELL J. M.,

BAXTER K. & CATHRO D. L. 1998. Structural evolution of the

Browse basin, North West Shelf: new concepts from deep-seismic

data. The sedimetary basins of Western Australia 2: Proceedings of

Petroleum Exploration Society of Australia Symposium, Perth,

WA, pp. 345–367. Petroleum Exploration Society of Australia.

SYMONDS P. A. & CAMERON P. J. 1977. The structure and stratigraphy

of the Carnarvon Terrace and Wallaby Plateau. APEA Journal

17, 30–41.

SYMONDS P. A., COLLINS C. D. N. & BRADSHAW J. 1994. Deep structure

of the Browse Basin: implications for basin development and

petroleum exploration. The Sedimentary Basins of Western

Australia. Proceedings of the Petroleum Exploration Society of

Australia, Perth 315–332.

TULLOCH A. J., BEGGS M., KULA J. L., SPELL T. L. & MORTIMER N. 2006.

Cordillera Zealandia, the Sisters Shear Zone and their influence

on the early development of the Great South Basin, New Zealand.

New Zealand Petroleum Conference, New Zealand. Ministry of

Economic Development.

WOODWARD N. B. 1995. Thrust systems in the Tamworth Zone,

southern New England Orogen, New South Wales. Australian

Journal of Earth Sciences 42, 107–117.

WOPFNER H. 1974. Post-Eocene History and Stratigraphy of North-

eastern South Australia. Transactions of the Royal Society of

South Australia 98, 1–12.

WOPFNER H. 1978. Silcretes of northern South Australia and adjacent

regions. In: Langford-Smith T. ed., Silcrete in Australia. pp. 93–

141, Department of Geography, University of New England,

Armidale.

ZOBACK M. 1992. First- and second-order patterns of stress in the

lithosphere: the World Stress Map Project. Journal of Geophysi-

cal Research 97, 11,703–11,728.

Received 2 February 2011; accepted 13 July 2011

1760

1765

1770

1775

1780

1785

1790

1795

1800

1805

1810

1815

1820

1825

1830

1835

1840

1845

1850

1855

1860

1865

1870

1875

1880

Australian paleo-stress 15

Page 16: Australian paleo-stress fields and tectonic …...strong geological plate elements and changes in far-field plate boundary forces have caused intraplate orogenesis and/or tectonic

1885

1890

1895

1900

1905

1910

1915

1920

1925

1930

1935

1940

1945

1950

1955

1960

1965

1970

1975

1980

1985

1990

1995

2000

2005

2010

APPENDIX 1 PROJECTION PARAMETERS

Listed here are the projections used for transforming

points of latitude and longitude along the plate margins

to and from a Cartesian reference frame for use in

ABAQUS.

CONTEMPORARY

Lambert conical conformal projection with an origin of

1328E longitude, 258S latitude and parallels at 158S and

358S

MIOCENE

Lambert conical conformal projection with an origin of

1308E longitude, 378S latitude and parallels at 08S and

408S

EOCENE

Lambert conical conformal projection with an origin of

132.58E longitude, 508S latitude and parallels at 208S and

508S

CRETACEOUS

Lambert conical conformal projection with an origin of

1208E longitude, 508S latitude and parallels at 408S and

608S.

16 R. D. Muller et al.